Skip to main content

Full text of "Dynamically-Driven Enhancement of the Catalytic Machinery of the SARS 3C-Like Protease by the S284-T285-I286/A Mutations on the Extra Domain."

See other formats


OPEN 3 ACCESS Freely available online 

Dynamically-Driven Enhancement of the Catalytic A 
Machinery of the SARS 3C-Like Protease by the S284- 
T285-I286/A Mutations on the Extra Domain 

Liangzhong Lim 1 ®, Jiahai Shi 19 ", Yuguang Mu 2 , Jianxing Song 1 * 

1 Department of Biological Sciences, Faculty of Science, National University of Singapore, Singapore, Republic of Singapore, 2 School of Biological Sciences, Nanyang 
Technological University, Singapore, Republic of Singapore 

Abstract 

Previously we revealed that the extra domain of SARS 3CLpro mediated the catalysis via different mechanisms. While the 
R298A mutation completely abolished the dimerization, thus resulting in the inactive catalytic machinery, N214A inactivated 
the enzyme by altering its dynamics without significantly perturbing its structure. Here we studied another mutant with 
S284-T285-I286 replaced by Ala (STI/A) with a 3.6-fold activity increase and slightly enhanced dimerization. We determined 
its crystal structure, which still adopts the dimeric structure almost identical to that of the wild-type (WT), except for slightly 
tighter packing between two extra-domains. We then conducted 100-ns molecular dynamics (MD) simulations for both STI/ 
A and WT, the longest reported so far for 3CLpro. In the simulations, two STI/A extra domains become further tightly 
packed, leading to a significant volume reduction of the nano-channel formed by residues from both catalytic and extra 
domains. The enhanced packing appears to slightly increase the dynamic stability of the N-finger and the first helix residues, 
which subsequently triggers the redistribution of dynamics over residues directly contacting them. This ultimately enhances 
the dynamical stability of the residues constituting the catalytic dyad and substrate-binding pockets. Further correlation 
analysis reveals that a global network of the correlated motions exists in the protease, whose components include all 
residues identified so far to be critical for the dimerization and catalysis. Most strikingly, the N214A mutation globally 
decouples this network while the STI/A mutation alters the correlation pattern. Together with previous results, the present 
study establishes that besides the classic structural allostery, the dynamic allostery also operates in the SARS 3CLpro, which 
is surprisingly able to relay the perturbations on the extra domain onto the catalytic machinery to manifest opposite 
catalytic effects. Our results thus imply a promising avenue to design specific inhibitors for 3CL proteases by disrupting their 
dynamic correlation network. 



@ PLOS | one 



Citation: Lim L, Shi J, Mu Y, Song J (2014) Dynamically-Driven Enhancement of the Catalytic Machinery of the SARS 3C-Like Protease by the S284-T285-I286/A 
Mutations on the Extra Domain. PLoS ONE 9(7): el 01 941. doi:10.1371/journal.pone.0101941 

Editor: Franca Fraternali, King's College, London, United Kingdom 

Received January 13, 2014; Accepted June 13, 2014; Published July 18, 2014 

Copyright: © 2014 Lim et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted 
use, distribution, and reproduction in any medium, provided the original author and source are credited. 

Funding: This study was supported by Ministry of Education of Singapore (MOE) Tier 3 Grant R-1 54-002-580-1 1 2, and also partly by Tier 2 Grant MOE 201 1 -T2-1 - 
096 (R1 54-000-525-1 12) to Jianxing Song. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the 
manuscript. 

Competing Interests: The authors have declared that no competing interests exist. 
* Email: dbssjx@nus.edu.sg 

9 These authors contributed equally to this work. 

n Current address: Whitehead Institute for Biomedical Research, Cambridge, Massachusetts, United States of America 



Introduction 

In 2002, severe acute respiratory syndrome (SARS) suddenly 
broke out in China and then rapidly spread to 32 countries, 
resulting in ~8500 infections and over 900 deaths (http:/ /www. 
who.int/csr/ sars/ en/). It is the first emerging infectious disease of 
the 21st century and was caused by a coronavirus termed SARS- 
CoV. Although now SARS appears to be contained, new 
coronaviruses have been detected, which may cause great threats 
to the human health. For example, since the appearance of a new 
coronavirus termed Middle East respiratory syndrome coronavirus 
(MERS-CoV) in April 2012, it has caused 207 confirmed cases, 
out of which 84 died (http://www.who.int/csr/disease/ 
coronavirus_infections/en/index.html). More importandy, so far 
neither a vaccine nor an efficacious therapy has been available for 
them. Therefore, it remains highly demanded to develop strategies 



to design potential therapeutic agents against SARS- and other 
CoVs. 

Among the known RNA viruses, coronaviruses are enveloped, 
positive-stranded ones with the largest single-stranded RNA 
genome (27-31 kilobases). The large replicase gene encodes two 
viral polyproteins, namely ppla (486 kDa) and pplab (790 kDa), 
which have to be processed into active subunits for genome 
replication and transcription by two viral proteases [1,2], namely 
the papain-like cysteine protease (PL2pro) and 3C-Like protease 
(3CLpro), also known as main protease (Mpro). Previously, SARS 
3CLpro has been extensively characterized to be a key target for 
development of antiviral therapies. The coronavirus 3CLpro is so 
named to reflect the similarity of its catalytic machinery to that of 
the picornavirus 3C proteases [1-3]. Noticeably, both 3C and 
3CL-Like proteases utilize the two-domain chymotrypsin fold to 
host the complete catalytic machinery, which is located in the cleft 



PLOS ONE | www.plosone.org 



1 



July 2014 | Volume 9 | Issue 7 | e101941 



Dynamical Enhancement of SARS-CoV 3CLpro 



between domains I and II. Intriguingly, however, in the 
coronavirus 3CLpro, a ~100-residue helical domain was evolu- 
tionarily acquired at its C-terminus [1-4]. Moreover, unlike 3C 
protease, only the homodimeric form is catalytically competent for 
the CoV 3CLpro [1,4,5-24]. After intense studies, now it has been 
clear that both chymotrypsin fold and extra domain are critical for 
dimerization. 

We were particularly interested in understanding the role of the 
extra domain and thus initiated a domain dissection study on 
SARS 3CLpro immediately after the SARS outbreak in Singapore 
[6] . The results revealed that although the catalytic fold and extra 
domain could fold independendy, the catalytic fold alone was 
monomeric and almost inactive. This indicates that the extra 
domain plays a key role in maintaining the dimerization, thus 
mediating the catalysis [6]. Therefore, we further conducted a 
systematic mutagenesis study which led to identification of the 
extra-domain residues critical for both dimerization and catalysis 
[16]. Interestingly, we found that the residues important for 
catalysis and dimerization constitute a nano-channel, which are 
composed of residues from both catalytic and extra domains [16]. 
Moreover, we determined the high-resolution structure of R298A, 
a monomeric mutant triggered by a point mutation on the extra 
domain, in which the most radical changes have been found 
within the catalytic machinery [22]. R298A adopts a completely 
collapsed and inactivated catalytic machinery which is structurally 
distinguishable from that in wild-type (WT) enzyme; with a short 
3 10 -helix formed by residues Serl39-Phel40-Leul41 within the 
oxyanion-binding loop [22]. Remarkably, the collapsed catalytic 
machinery observed in R298A appears to represent a universal 
inactivated state intrinsic to all inactive enzymes as the same 
collapsed machinery was found in other monomers triggered by 
the mutations G11A, N28A and S139A which are all located on 
the chymotrypsin fold [21,23,24]. On the other hand, previously 
we also identified a mutant N214A, which owns a dramatically 
abolished activity but only slightly weakened dimerization [16]. 
Very unexpectedly, our determination of its crystal structure 
revealed that it still adopts a dimeric structure almost identical to 
that of the WT protease [25]. Nevertheless, the results with 
molecular dynamics (MD) simulations up to 30 ns unveiled that 
the N214A mutation triggers the dynamical instability of the 
catalytic machinery, with many key residues jumped to sample the 
conformations characteristic of the collapsed and inactivated state 
in R298A, thus establishing a dynamically-driven inactivation 
mechanism for the catalytic machinery of the SARS 3CLpro [25]. 

Here we studied another mutant we previously identified with 
three residues S284-T285-I286 mutated to Ala, designated as 
STI/A [16]. Interestingly, despite being far away from the active 
pocket, the mutations led to a 3.6-fold enhancement of the 
catalytic activity but only slighdy enhanced dimerization. Here, 
our determination of its crystal structure reveals that STI/ A still 
adopts the dimeric structure almost identical to that of the wild- 
type (WT), only with a slighdy change of the extra-domain 
packing, which is similar to that observed in the more active form 
of the WT at high pH values (pH 7.6 and 8) [44]. As a 
consequence, to understand its underlying dynamical mechanism, 
we conducted 100-ns MD simulations for both WT and STI/A. 
Remarkably, the most dramatic changes in STI/A simulations are 
associated with the nano-channel. This change appears to slighdy 
enhance the dynamic stability of the N-finger and helix A, which 
subsequently relay the dynamic effects to the contacted residues 
and finally to catalytic machinery. Ultimately, the key components 
composed of the catalytic machinery become more dynamically 
stable, thus rationalizing the enhancement of its catalytic activity. 
As coronavirus 3C-Like proteases share a similar enzymatic 



mechanism [1,2,26], our results would facilitate the development 
of strategies and agents by modulating protein dynamics to fight 
new coronaviruses whose outbreaks may occur in the future. 

Materials and Methods 

Accession Numbers 

The structure coordinate of the STI/ A mutant was deposited in 
Protein Data Bank, with PDB ID code of 3EA8. 

Generation of the recombinant STI/A mutant without 
any extra residue 

Recently the extra N-terminal residues leftover from the 
cleavage of fusion proteins were demonstrated to significandy 
reduce the enzymatic activity [10,18]. On the other hand, the 
SARS 3CLpro we previously studied had two extra residues Gly- 
Ser after the thrombin cleavage of the GST-3CLp fusion proteins 
[6,16]. In order to remove the two extra residues, in the present 
study we transferred the gene encoding SARS 3CLp from the 
pGEX-4T- 1 GST-fusion expression vector (Amersham Bioscienc- 
es, GE Healthcare, Little Chalfont, UK) to the His-tagged pET28a 
vector. Subsequendy, site-directed mutagenesis was utilized to 
shorten the thrombin cleavage sequence LVPR | GS (CTG GTT 
CCG CGT GGA TCC) engineered by the company to LVPR| 
(CTG GTT CCG CGT), which only constituted the thrombin 
cleavage site in conjunction with the first two N-terminal residues 
Ser-Gly of SARS 3CLp. Interestingly, thrombin cleaved this new 
site (LVPR | SG) very efficiently to release the authentic wild-type 
3CLpro as well as N214A [25]. To produce STI/A mutant, site- 
directed mutagenesis was further used to mutate S284-T285-I286 
to Ala as previously described [16,25]. Benefited from this re- 
engineered cleavage site, we were able to produce both authentic 
WT 3CLpro [25] and its STI/A mutant without any extra 
residues from the fusion tag. 

Experimentally, the His-tagged protease or mutants was 
expressed in E. coli strain BL21 (DE3) with induction by 
0.4 mM isopropyl-l-thio-d-galactopyranoside (IPTG) under 20 
degree overnight. The protease was obtained by in-gel cleavage 
with thrombin while the His-tag protein attached to the Nickel- 
NTA beads (QIAGEN), foUowed by a FPLC purification on a gel 
filtration column (HiLoad 16/60 Superdex 200). The pure 
protease was concentrated to 10 mg/ml and stored in the buffer 
(10 mM Tris-HCl, 10 mM DTT, pH 7.5). The molecular weight 
of the WT and N2 1 4A proteases were determined by a Voyager 
STR MALDI-TOF mass spectrometer (Applied Biosystems). 

Enzymatic activity assay and ITC characterization of 
dimerization 

The enzymatic activities of the STI/A proteases were 
measured by a fluorescence resonance energy transfer (FRET)- 
based assay using a fluorogenic substrate peptide (Dabcyl- 
KTSAVLQSGFRKME-Edans) as previously described 
[15,16,25,27]. Briefly, 1 ml reaction mixture contained 50 nM 
protease and fluorogenic substrate with concentrations ranging 
from 1 uM to 40 uM in a 5 mM Tris-HCl buffer with 5 mM 
DTT at pH 6.0, which is identical to the pH for crystallization. 
The enzyme activity was measured by monitoring the increase of 
the emission fluorescence at a wavelength of 538 nm with 
excitation at 355 nm using a Perkin-Elmer LS-50B luminescence 
spectrometer. The Km and kcat values were deduced from data 
analysis using Graphpad prism. 

ITC (isothermal titration calorimetry) experiments were carried 
out to determine the monomer-dimer dissociation constants of the 
STI/A proteases as previously described [10,25] using a Microcal 



PLOS ONE | www.plosone.org 



2 



July 2014 | Volume 9 | Issue 7 | e101941 



Dynamical Enhancement of SARS-CoV 3CLpro 



VP ITC machine. Briefly, the protease samples and buffers were 
span at 13.3k rpm for one hour to remove the tiny particles and 
degas thoroughly. In titrations, the STI/A sample in 5 mM Tris- 
HC1 buffer at pH 6.0 containing 5 mM DTT were loaded in the 
syringe, which was subsequently titrated into the same buffer in 
the cell. The obtained titration data with endothermic peaks were 
analyzed by the built-in Microcal ORIGIN software using a 
dimer-monomer dissociation model to generate the dissociation 
constants and the enthalpy changes. 

Crystallization, structure determination of the STI/A 
protease 

The STI/A protease with a concentration of 10 mg/ml were 
crystallized in a 2 (0.1 hanging drop using a condition identical to 
that previously reported except for a minor variation of the 
PEG6000 concentration [4,18,25]. After a three-day growth, a 
single crystal was picked up from the crystal clusters for diffraction 
with the cryoprotective buffer (20% glycerol with the mother 
liquid). The X-ray diffraction data were collected at Bruker X8 
PROTEUM in-house X-ray machine. 

The collected data sets were processed using the program 
HKL2000 in a resolution of 2.25 A in the space group C2. Briefly, 
the phase determination for the mutant structure was done by the 
molecular replacement method using the WT SARS-CoV 3CLpro 
structure (PDB code: 2H2Z) [18] as the searching model by the 
program Phaser [28] in the program suite Phenfx [29]. The Ala 
mutating residues (Ser284, Thr285, Ile286) were corrected in the 
program COOT [30] . The refinements and the addition of the 
solvent molecules of the models for the mutant were done in the 
program suite Phenix. The final model was analyzed by 
PROCHECK [31] and all figures were prepared using Pymol 
[32]. 

Molecular dynamics (MD) simulations 

The crystal structures of the WT (PDB code: 2H2Z) [18] and 
STI/A determined in the present study were selected as the initial 
models for molecular dynamics simulations as previously described 
[25]. The crystal structures PDB files were post-processed as 
previously described [33,34]. 

The simulation cell is a periodic cubic box with a minimum 
distance of 10 A between the protein and the box walls to ensure 
the protein would not directiy interact with its own periodic 
images given the cutoff. The water molecules, described using the 
TIP3P model, were filled in the periodic cubic box for the all atom 
simulation, 6 Na + ions were randomly placed to neutralize the 
charge in MD system for STI/A and WT dimers, Each system 
contained approximately 75,000 atoms. 

Three independent 100-ns MD simulations for each protease 
were performed with the program GROMAGS [35] with the 
AMBER-03 [36] all-atom force field. The long-range electrostatic 
interactions were treated using the fast particle-mesh Ewald 
summation method [35], with the real space cutoff of 9 A and a 
cutoff of 14 A was used for the calculation of van der Waals 
interactions. The temperature during the simulations was kept 
constant at 300 K by Berendsen's coupling. The pressure was held 
at 1 bar. The isothermal compressibility was 4.5*10~ 5 bar" 1 . The 
time step was set as 2 fs. All bond lengths including hydrogen 
atoms were constrained by the LINCS algorithm [37]. Prior to 
MD simulations, all the initial structures were relaxed by 500 steps 
of energy minimization using the steepest descent algorithm, 
followed by 100 ps equilibration with a harmonic restraint 
potential applied to all the heavy atoms of the protease. 



Calculation of enclosed volume 

POVME program was used to calculate the enclosed volumes of 
the nano-channel and Thr25-Cys44 Leu-P2 substrate pocket [38]. 
Snapshots from each MD trajectory were taken at every 100-ps 
interval. A 3D-grid spacing was constructed around the C p of 
residues: Ser284, Thr285, Ile286 in the nano-channel, and Thr25 
and Cys44 in the Leu-P2 substrate pocket. Grid points were 
included when they do not overlap with any atom of the studied 
residues or neighboring residues. The grid points were summed up 
to give the enclosed volume. 

Calculation of center of mass (COM), angle © 

Heavy atoms of all atoms are used in the computation of center 
of mass: C, N, O and S. Hydrogen atoms are ignored. The COM 
is calculated at lps interval. For the angle 0 calculation, COM of 
the two chymotrypsin fold (residues 7-180) of both protomers and 
the extra-domain/Domainlll (residue 200-302) of each protomer. 
A vector is computed for each COM of Domain III to the COM 
of the 2 chymotrypsin-folds. The angle © is the angle between 
these 2 vectors. 

Correlation analysis 

STI/A as well as N214A modulates the catalysis of the SASR 
3CL protease without substantial conformational change, suggest- 
ing that the mutation effects are relayed to the catalytic machinery 
through dynamic allostery. So here we attempted to analyze the 
allosteric networks of WT, N214A and STI/A by a recendy 
established approach called Mutlnf [39]. Mutlnf represents an 
entropy-based approach to analyze ensembles of protein con- 
formers, such as those from molecular dynamics simulations by 
using internal coordinates and focusing on dihedral angles. In 
particular, this approach is even applicable in cases for which 
conformational changes are subtie, for example, because the 
coupling is mostiy entropic in nature [39]. Briefly, this approach 
utilizes second-order terms from the configurational entropy 
expansion, called the mutual information, to identify pairs of 
residues with correlated conformations, or correlated motions, in 
an equilibrium ensemble [39]. In the present study, the 
normalized matrix values were used, and 0.3 was set up to be 
the threshold value to determine the pairs of highly correlated 
residues. The residue pairs with correlated values &0.3 make up 
the top 1% of the entire matrix. 

Results 

Activity and dimerization of the STI/A enzyme 

By reengineering the thrombin cleavage site, we succeeded in 
obtaining the STI/A proteases without any extra residues leftover 
from the fusion tag. The enzyme was characterized by far-UV CD 
spectroscopy and its spectrum (spectra not shown) had no 
detectable difference from that with two extra residues Gly-Ser 
that we previously studied [16], thus indicating that the two extra 
residues have no detectable effect on the secondary structures. 

By a fluorescence resonance energy transfer (FRET)-based 
assay, we have measured the enzymatic activities of the authentic 
STI/A proteases. As shown in Figure la, the STI/A protease has 
a Km value (23.3 uM) very similar to those previously reported on 
the authentic wild-type enzyme (24.2 p.M) while Kcat of the STI/ 
A protease has a 3.5-fold enhancements [25]. Here we further 
measured the Kd value of the STI/A dimerization by ITC to be 
13.4 uM (Figure lb), only slightly smaller than 21.4 uM for the 
authentic WT protease [25], indicating that the STI/A mutation 
only has a very small enhancement of the dimerization. 



PLOS ONE | www.plosone.org 



3 



July 2014 | Volume 9 | Issue 7 | e101941 



Dynamical Enhancement of SARS-CoV 3CLpro 




Equivalent monomer concentration 
(mM) 

Figure 1. Enzymatic activity and dissociation constant of the dimer-monomer equilibrium, (a). Enzymatic activity of STI/A vs. substrate 
concentrations as measured by monitoring the increase of the emission fluorescence intensity at a wavelength of 538 nm. The Km and kcat values 
are presented, (b). The ITC dilution profiles for measuring the dissociation constant of the dimer-monomer equilibrium for STI/A. The Kd and AH 
values were obtained by fitting the ITC data with the built-in Microcal ORIGIN software, (c). Dimeric structure of the SARS 3CLpro with the catalytic 
folds colored in blue (protomer 1) and cyan (protomer 2); the extra domains in red (protomer 1) and lightpink (protomer 2), as well as the connecting 
loops in black (protomer 1) and purple (protomer 2). The catalytic dyad residue His41 is displayed as green sphere while Cys145 as yellow sphere. 
Mutation residues S284T285I286 and N214 are displayed as spheres and the N-finger residues Ser1-Fhe8 are displayed as dots. 
doi:1 0.1 371 /journal.pone.01 01 941 .g001 



Crystal structure of the STI/A enzyme 

To gain insights into the structural consequence of the STI/A 
mutation, we determined its crystal structure at a resolution of 
2.25 A in C2 space group, with one molecule in the asymmetric 
unit. The R wor k factor of the final models for STI/A mutant was 
18.4%, while the R fr( . e factor was 23.1%. Details of the data 
collection and refinement statistics are presented in Table SI. 

In the electron density map of the STI/A mutant, all 306 
residues are visible and it adopts the classic dimeric structure with 
the same packing of the two protomers as observed in all 
previously-determined dimeric structures of the coronavirus 
3CLpro [1,2,4,18,19,25,40-45]. The dimeric STI/A structure is 
highly similar to those of the WT protease previously reported 
(Figure 2a). If compared to the WT crystal structure (2H2Z) [18] 
also determined at pH 6.0, with the authentic sequence in the 
same C2 space group [18], the dimeric RMS deviations are 0.86 
and 0.77 A respectively for the heavy (C, O, S and N) and 
backbone atoms (C, Ca, N (backbone-amide), carbonyl). Even if 
comparing at the individual protomer level, the RMS deviations 
reduce to 0.55 and 0.49 A respectively for the heavy and backbone 
atoms, implying that the packing of two protomers slightly differs 



in the STI/A proteases. In the STI/A mutant, even for the key 
residues constituting the catalytic machinery including the 
catalytic dyad His41-Cysl45, oxyanion-loop Phel40-Cysl45, 
Hisl63 and Glul66 critical for binding substrates; and Phel40, 
His 172 in holding the substrate binding-pocket open, their 
backbones and side-chains are almost superimposable to those in 
the WT structure (Figure 2b). This observation implies that the 
enhancement of the STI/A activity cannot be readily rationalized 
only by the static structure. Furthermore, the mutation residues 
Ser284-Thr285-Ile286 are located on the extra domain and far 
away from the catalytic dyad His41-Cysl45 (Figure lc). We also 
examined the B-factors of STI/A and WT proteases but it appears 
extremely challenging to establish any precise correlation between 
the B-factors and catalytic activity. Therefore, the catalytic 
enhancement is most likely due to the change of the protein 
dynamics of the enzyme induced by the mutations as we 
previously demonstrated on the N214A mutant [25]. 

Intriguingly, the enzymatic activity of the wild-type 3CLpro has 
been previously found to be pH-dependent where the optimum 
activity is at pH 7.6 and the activity become much lower at 
pH 6.0 [4,5,18,44—45]. Strikingly, upon pH changes, the extra 



PLOS ONE | www.plosone.org 



4 



July 2014 | Volume 9 | Issue 7 | e101941 



Dynamical Enhancement of SARS-CoV 3CLpro 



Catalytic^ 
Dyad V 




Figure 2. Crystal structure of the STI/A mutant, (a). Overall superimposition of the dimeric STI/A (violet) and WT (cyan; PDB code of 2H2Z (18) 
structures, (b) Superimpositions of the catalytically critical residues of STI/A (violet) and WT (cyan; PDB code of 2H2Z (18). (c). Diagram showing the 
distance (d) between the mass centers of two extra domains; and the angle (0) by the mass centers of two extra domains as well as the mass center 
of two chymotrypsin folds together (see Material and Methods for more details). The cavity volumes (V) of the nano-channels of STI/A (d) and WT (e) 
as represented by violet dots, which were calculated with the program: POVME (38). 
doi:1 0.1 371 /journal.pone.01 01 941 .g002 



domains in the dimeric enzyme appear to undergo a 'rigid body 
rotation/movement'. Inspired by this, here we calculated the 
distances (d) between the centers of mass of two extra domains, as 
well as the angles (©) formed by two mass centers of each extra 
domains and the mass center of the two catalytic folds together 
(Figure 2c), which are 32.4 A and 63.2 degree for the STI/A 
determined here at pH 6.0 (3EA7), 33.7 A and 66.3 degree for the 
wild-type enzyme at pH 6.0 (2H2Z) [18], 32.6 A and 63.3 degree 
for the wild-type enzyme at pH 7.6 (1UK3), and 32.7 A and 63.7 
degree for the wild-type enzyme at pH 8.0 (1UK2). This 
observation implies that the slightly tighter packing of the two 
extra domains is indeed associated with the higher catalytic 
activity. 

Previously, based on a systematic mutagenesis study, we have 
proposed a nano-channel to be responsible for relaying the 
mutation effect from the extra domain to the catalytic machinery, 
which are mainly constituted by residues next to S284-T285-I286 
on the extra domains as well as N-ffnger residues on the catalytic 



folds [16]. Because the mutation of STI to Ala with a small side 
chain results in a cavity, in STI/A the two extra domains come 
slightly closer as described above, and consequently the volume of 
the nano-channel (V) is 1293 A 3 in STI/A (Figure 2d), slightly 
smaUer than that (1435 A 3 ) in WT (Figure 2e). 

Molecular dynamics (MD) simulations 

Molecular dynamics simulation is a powerful tool to provide 
insights into the dynamic mechanism that underlies protein 
function [25,46-50]. Therefore, we conducted 100-ns MD 
simulations for the wild-type (2H2Z) [18] and our present STI/ 
A (3EA7) structures, both of which have been determined at 
pH 6.0 and in crystals of the space group C2. 

Figure 3 presents the root-mean-square deviation (RMSD) of 
Cot atoms (from their positions in the energy minimized structures) 
for the dimeric WT and STI/A (a-c), their catalytic folds (domains 
I and II; d-f) and extra domains only (domain III, g-j). 
Interestingly, in the context of the dimeric forms, STI/ A showed 



PLOS ONE | www.plosone.org 



5 



July 2014 | Volume 9 | Issue 7 | e101941 



Dynamical Enhancement of SARS-CoV 3CLpro 



larger overall RMSD than WT, with average RMSD values over 
100-ns simulations of 2.64±0.32, 1.90±0.26 and 2.68±0.42 A for 
three separate simulations of STI/A; and 1.60±0.26, 1.66±0.31 
and 1.63±0.25 A for three simulations of WT. However, the 
averaged Cot RMSD values of the chymotrypsin folds (Domain I 
and II) of STI/A are only slightly larger than those of WT 
(Figures 3d-3f), with average RMSD values over 100-ns simula- 
tions of 1.40±0.21, 1.20±0.16 and 1.44±0.26 A for three 
simulations of STI/A; and of 1.28±0.16, 1.17+0.17 and 
1 .25 ±0.15 A for three simulations of WT. On the other hand, 
the extra domains of STI/A have dramatically higher fluctuations 
than WT in simulations as indicated by its higher RMSD values 
(Figures 3g-3i), with average RMSD values over 100-ns simula- 
tions of 3.57±0.54, 2.19±0.53 and 3.26±0.86 A for three 
simulations of STI/A and of 1.75±0.51, 1.62±0.31 and 
1.61 ±0.31 A for three simulations of WT. These results clearly 
demonstrate that the larger RMSD values for the dimeric STI/A 
are mostiy resulting from the motions of the extra domains. This 
premise is further evident from the root-mean-square fluctuation 



(RMSF) averaged over 100 ns for both STI/A and WT residues 
(Figures 3j-3o). 

Indeed, for STI/A, the distance d (Figures 4a-c) and angle © 
(Figure 4d-f) as defined in Figure 2d have further reduced after 
~10 ns in simulations 1 and 3 and after ~45 ns in simulation 2. 
For the average distance over 100-ns trajectories, STI/A has 
28.77±0.90, 30.69± 1.16 and 29.20± 1.29 A respectively for three 
simulations, while WT has 32.78±1.07, 34.12±0.62 and 
33.33±0.656 A respectively. For the average angle over 100-ns 
trajectories, STI/A has 54.78± 1.781, 58.98±2.544 and 
55.25±2.535 degree respectively for three simulations, while 
WT has 64.54± 2.250, 66. 65 ±1.461 and 65.43 ±1.614 degree 
respectively. These results indicate that in all three separate 
simulations, the two extra domains of STI/A become more tightly 
packed than those of WT. Consequently the volumes of the nano- 
channel significantly reduced for STI/ A in all three simulations as 
compared to those of WT (Figures 4g-h), with average values of 
512.62±212.52, 890.26±247.17 and 579.80±261.51 A 3 respec- 
tively for three simulations of STI/A, and of 1204.35±255.69, 
1428.93± 121.50 and 1319.94± 143.21 A' 5 respectively for WT. 



f (b) M (c) ] 




_l i i i i I i i i i L 



_l i i i i I i i i i L 



Residue 

Figure 3. Overall dynamic behaviors in the W1D simulations. Root-mean-square deviations (RMSD) of the Cot atoms (from their positions in the 
energy minimized structures) for three independent MD simulations of the dimeric STI/A (black) and WT (red) for the whole enzyme (a-c); catalytic 
fold (d-f); and extra domain (g-i). Root-mean-square fluctuations of the Ca atoms computed for three independent simulations for protomer A (j-l) 
and protomer B (m-o) of STI/A (black) and WT (red). Protomer A and B are denoted as PI and P2 respectively. 
doi:1 0.1 371 /journal.pone.01 01 941 .g003 



PLOS ONE | www.plosone.org 



6 



July 2014 | Volume 9 | Issue 7 | e101941 



Dynamical Enhancement of SARS-CoV 3CLpro 




Time (ns) 



Figure 4. Changes of the packing interaction between two extra domains. Three separate time-trajectories of the distance d (a-c); angle 0 
(see Figure 2d) (d-f) and volume of nano-channel (residue 284-286), V (g-i) of STI/A (black) and WT (red). 
doi:1 0.1 371 /journal.pone.01 01 941 .g004 



We have also calculated the hydrogen bond occupancy of all 
simulations for both STI/A and WT, and Table 1 presents the 
hydrogen bond occupancies with the difference >5% between 
STI/A and WT. Interestingly, although the STI/A mutations are 
on the extra domain, only limited intra-domain hydrogen bonds 
show significant occupancy differences within the extra domain 
(Table 1). This suggests that similar to the tighter packing between 
two extra domains triggered by the high pH in WT [44] , here the 
tighter packing induced by the STI/A mutations is also largely 
resulting from the 'rigid body rotation/movement' of the extra 
domains without significantly affecting the packing within the 
extra domains. Furthermore, the slight dynamic changes over the 
extra domains can also be transmitted to the catalytic fold, as 
exemplified by the significant changes of hydrogen bond 
occupancies between Asp295 and Thrill (Table 1). 

Dynamic behavior of the catalytic dyad and oxyanion 
hole 

In the previously-determined crystal structures of SARS 
3CLpro, the distance between NE2 of His41 and SG of Cysl45 
ranges from 3.6 to 3.9 A. Furthermore, all previous MD 
simulations revealed that the dynamic stability of this distance is 
extremely critical for the stable formation of a hydrogen bond, 
which appears to be pivotal for maintaining the catalytic 
competency of the SARS 3CLpro [10,23,25,51-53]. Also for the 
active WT enzyme, this distance has been previously demonstrat- 
ed to be dynamically stable in MD simulations. Indeed, as in our 
previous simulations, the distances of both protomers of the WT 
were found to be dynamically stable for all 30 ns while those of the 
inactive N214A mutant became unstable and larger despite 
sharing a superimposable crystal structure to the WT [25]. In our 
current MD simulations of 100 ns, the distances between NE2 of 
His41 and SG of Cysl45 of STI/A are: 3.80±0.60 A, 3.91±0.51 
and 3.78±0.50 A respectively for three simulations of the 
protomer 1 (Figures 5a-5c), and 3.91 ±0.48, 3.87±0.45 and 
3.86±0.48 A respectively for the protomer 2 (Figures 5d-5f). In 
contrast, the distances in WT are: 3.82±0.46, 3.83±0.49 and 



4.07 ±0.76 A respectively for three simulations of the protomer 1, 
and 4.08±0.52, 3.84±0.49 and 3.89±0.43 A respectively for the 
protomer 2. As a consequence, the distance of STI/A ranges from 
3.78 to 3.91 A while WT from 3.82 to 4.08 A, suggesting that 
overall this distance is more dynamically stable in STI than that in 
WT. In particular, after 80 ns, in simulation 3 of the WT 
protomer 1, this distance becomes largely fluctuating (Figure 5c), 
while for STI/A, no such large fluctuation can be observed for all 
100 ns. The Chi2 angles of His41 have similar dynamical 
behaviors (Figures 5g-51) and the large fluctuation is also observed 
(Figure 5i). 

One key component of the catalytic machinery of the SARS 
3CLp is the substrate-binding pocket composed of six substrate 
sites, namely S1-S6, which correspond to the P1-P6 residues of 
the substrate [1,2,4,23,25]. The SI substrate site is the most critical 
as it confers an absolute specificity for a Gin at the PI position of 
the substrate. As such, maintaining the intact conformation of SI 
substrate site is especially vital for catalysis. Briefly, the SI 
substrate site can be divided into four parts: the oxyanion hole; 
Hisl63; Glul66; Phel40 and its stabilizing elements [1,2,4,23,25]. 
The oxyanion hole refers to a structural element to donate two 
hydrogen bonds from main-chain amides of Gly- 1 43 and Cys- 145 
to accommodate the main-chain oxygen of Gln-P 1 as well as the 
tetrahedral intermediate during catalysis. Previously we demon- 
strated that in the inactive monomer R298A, the most 
distinguishable change is the collapse of the oxyanion hole as 
triggered by the chameleon formation of a short 3 10 -helix from a 
loop over residues Serl39-Phel40-Leul41 [22]. Remarkably, in 
our previous MD simulations, residues Ser 1 39-Phe 1 40-Leu 1 4 1 of 
the WT maintained an extended active conformation while those 
of N2 1 4A frequently jumped to sample the 3 1 0 -helix conformation 
characteristic of the collapsed oxyanion hole [25] . 

In the present simulations of both STA and WT up to 100 ns, 
the three residues have similar dynamical behaviors in their 
backbone conformations (Figure SI), as well as the similar 
hydrophobic stacking interactions among Phel40, His 163 and 
His 172. 



PLOS ONE | www.plosone.org 



7 



July 2014 | Volume 9 | Issue 7 | e101941 



Dynamical Enhancement of SARS-CoV 3CLpro 



Table 1. Hydrogen Bond Occupancy for STI/A and WT with Significant Differences. 





Donor 








Acceptor 








Average (%) 








Res No. 


Name 


Atom 


Chain 


Res No. 


Name 


Atom 


Chain 


STI/A 


WT 


(STI-WT) 




Extra Domain 


276 


MET 


N 


A 


271 


LEU 


O 


A 


0.0 


5.5 


-5.5 




276 


MET 


N 


B 


271 


LEU 


O 


B 


0.2 


10.7 


-10.5 




278 


GLY 


N 


B 


285 


THR 


0 


B 


0.0 


13.9 


-13.9 




230 


PHE 


N 


A 


226 


THR 


0 


A 


62.4 


78.7 


-16.3 




230 


PHE 


N 


B 


226 


THR 


0 


B 


67.1 


82.5 


-15.4 




257 


THR 


N 


B 


253 


LEU 


0 


B 


47.1 


59.3 


-12.2 




299 


GLN 


N 


A 


296 


VAL 


0 


A 


13.1 


36.1 


-23 




299 


GLN 


N 


B 


296 


VAL 


0 


B 


29.1 


40.0 


-10.9 




-13.5 + 5.1 


Extra Domain-Chymotrypsin Fold 


111 


THR 


OG1 


A 


295 


ASP 


OD1 


A 


30.8 


0.1 


30.7 




111 


THR 


OG1 


B 


295 


ASP 


OD1 


B 


34.3 


15.7 


18.6 




24.7 ±8.6 


N-Fingers 


4 


ARG 


NH1 


A 


290 


GLU 


OE2 


B 


29.3 


44.6 


-15.3 




4 


ARG 


NH1 


B 


290 


GLU 


OE2 


A 


31.0 


54.6 


-23.6 




4 


ARG 


NH2 


A 


137 


LYS 


O 


B 


14.1 


24.4 


-10.3 




4 


ARG 


NH2 


B 


137 


LYS 


O 


A 


5.8 


22.9 


-17.1 




5 


LYS 


NZ 


A 


288 


GLU 


OE2 


A 


31.6 


20.6 


11.9 




5 


LYS 


NZ 


B 


288 


GLU 


OE2 


B 


21.3 


14.0 


7.3 




7 


ALA 


N 


A 


125 


VAL 


O 


B 


83.6 


67.1 


16.5 




-4.5+15.8 


Helix A at Domain 1 


10 


SER 


N 


A 


10 


SER 


OG 


B 


71.3 


34.3 


37.0 




10 


SER 


N 


B 


10 


SER 


OG 


A 


61.5 


33.8 


27.7 




11 


GLY 


N 


B 


14 


GLU 


OE1 


A 


93.3 


32.8 


60.5 




11 


GLY 


N 


A 


14 


GLU 


OE1 


B 


62.4 


72.8 


-10.4 




13 


VAL 


N 


A 


10 


SER 


O 


A 


87.6 


48.1 


39.5 




15 


GLY 


N 


B 


11 


GLY 


O 


B 


53.2 


9.1 


44.1 




95 


ASN 


ND2 


A 


15 


GLY 


O 


A 


72.7 


67.0 


5.7 




95 


ASN 


ND2 


B 


15 


GLY 


O 


B 


75.6 


63.9 


11.7 




27.0±23.2 


Asn28 


28 


ASN 


ND2 


A 


117 


CYS 


O 


A 


27.5 


10.4 


17.1 




28 


ASN 


ND2 


B 


117 


CYS 


O 


B 


32.4 


18.0 


14.4 




28 


ASN 


ND2 


A 


145 


CYS 


O 


A 


87.6 


30.5 


57.1 




28 


ASN 


ND2 


B 


145 


CYS 


O 


B 


78.0 


56.0 


22.0 




27.7±19.9 


Glu166 


172 


HIE 


NE2 


A 


166 


GLU 


OE2 


A 


43.3 


19.6 


23.7 




Linker between Catalytic and Extra Domains 


175 


THR 


OG1 


A 


176 


ASP 


O 


A 


16.9 


26.5 


-9.6 




175 


THR 


OG1 


B 


176 


ASP 


O 


B 


27.2 


39.2 


-12.0 




105 


ARG 


NH1 


A 


180 


LYS 


O 


A 


14.9 


25.1 


-10.2 




105 


ARG 


NH1 


B 


180 


LYS 


O 


B 


12.1 


34.3 


-22.2 




182 


TYR 


N 


A 


174 


GLY 


O 


A 


88.9 


96.4 


-7.5 




CO 


TYR 


N 


B 


174 


GLY 


O 


B 


89.4 


95.6 


-6.2 




-11.3 + 5.7 



PLOS ONE | www.plosone.org 



8 



July 2014 | Volume 9 | Issue 7 | e101941 



Dynamical Enhancement of SARS-CoV 3CLpro 



Table 1. Cont. 





Donor 








Acceptor 








Average (%) 






Res No. 


Name 


Atom 


Chain 


Res No. 


Name 


Atom 


Chain 


STI/A 


WT 


(STI-WT) 


Chymotrypsin Fold 


19 


GLN 


NE2 


A 


26 


THR 


OG1 


A 


36.3 


21.5 


14.8 


19 


GLN 


NE2 


B 


26 


THR 


OG1 


B 


59.5 


0.0 


59.5 


19 


GLN 


NE2 


A 


119 


ASN 


0 


A 


5.5 


20.3 


-14.8 


19 


GLN 


NE2 


B 


119 


ASN 


0 


B 


2.1 


11.3 


-9.2 


21 


THR 


OG1 


A 


25 


THR 


0 


A 


0.0 


8.8 


-8.8 


21 


THR 


OG1 


B 


25 


THR 


0 


B 


0.0 


18.2 


-18.2 


22 


CYS 


N 


A 


25 


THR 


0 


A 


98.2 


65.6 


32.6 


22 


CYS 


N 


B 


25 


THR 


0 


B 


98.5 


63.1 


35.4 


25 


THR 


OG1 


A 


44 


CYS 


0 


A 


85.9 


40.8 


45.1 


25 


THR 


OG1 


B 


44 


CYS 


0 


B 


75.9 


52.8 


23.1 


25 


THR 


N 


A 


22 


CYS 


0 


A 


47.8 


34.6 


13.2 


25 


THR 


N 


B 


22 


CYS 


0 


B 


43.4 


24.4 


19.0 


26 


THR 


OG1 


A 


21 


THR 


OG1 


A 


74.4 


62.3 


12,1 


26 


THR 


OG1 


B 


21 


THR 


OG1 


B 


85.0 


25.8 


59/2 


113 


SER 


OG 


A 


127 


GLN 


OE1 


A 


95.5 


84.4 


11.1 


113 


SER 


OG 


B 


127 


GLN 


OE1 


B 


92.7 


46.1 


46.6 


116 


ALA 


N 


A 


124 


GLY 


0 


A 


92.7 


75.6 


17.1 


117 


CYS 


N 


B 


147 


SER 


OG 


B 


80.2 


65.8 


14.4 


121 


SER 


N 


A 


118 


TYR 


0 


A 


65.1 


49.1 


16.0 


121 


SER 


N 


B 


118 


TYR 


0 


B 


61.6 


56.0 


5.6 


135 


THR 


N 


A 


133 


ASN 


OD1 


A 


63.6 


55.5 


8.1 


135 


THR 


N 


B 


133 


ASN 


OD1 


B 


61.6 


54.7 


6.9 


148 


VAL 


N 


A 


115 


LEU 


0 


A 


63.0 


20.8 


42.2 


-18.7±21.8 



doi:1 0.1 371 /journal.pone.01 01 941 .t001 



Dynamic behavior of Hisl 63 and Glu166 

The Glu-Pl substrate binding site includes essential substrate 
recognition residues: Hisl63, Glul66 and Hisl72. Extensive 
studies previously revealed that increased side-chain flexibility of 
His 163 and Glul66 decreases the enzymatic activity [44], and 
mutation of Glul66 to alanine reduces the enzymatic activity by 
many folds [54]. Particularly these three residues were found to 
play a key role in the pH-dependent enhancement of the catalytic 
activity by both experiments and simulations [44]: the reduced 
contact between His 163 and Glul66 allowed Glul66 to engage in 
hydrogen bonding with Hisl 72 and consequently adopted optimal 
conformations for substrate-binding at pH 7.6. In contrast, the 
increased contact between His 163 and Glul66 decreased hydro- 
gen bond between Glul66 and Hisl 72, and thus reduced the 
stability of the substrate pocket at pH 6 [44] . 

In our current 100-ns MD simulation, the backbone confor- 
mations of both His 163 and Glul66 are indistinguishable between 
STI/A and WT. In STI/A, the distances of Hisl63-Glul66 are: 
7.97±0.32, 7.75±0.35 and 7.89±0.35 A (with the average of 
7.87 A) for three simulations of protomer 1; 7.60±0.56, 
8.31 ±0.31 and 8.18±0.50 A (with the average of 8.03 A) for 
protomer 2. In WT, the distances are: 8.02±0.42, 7.94±0.37 and 
8.17±0.47 A (with the average of 7.86 A) for three simulations of 
protomer 1; 7.72±0.56, 8.05±0.34 and 7.81±0.44 A (with the 
average of 8.04 A) for protomer 2. Furthermore, the dynamic 



behavior of the Hisl63-Hisl 72 distance is highly similar in both 
STI/A and WT (Figure S2a-S2f). 

Interestingly, STI/ A showed enhanced dynamic stability in the 
Glul66 side-chain conformation (Figures 6a-6f) and smaller 
centroid distance between the side chains of Glul66 and His 172 
(Figures 6g-61). The averaged centroid distances of Glul66 and 
His 172 are: 4.80 A and 4.81 A respectively for two protomers of 
STI/A; and 4.73 and 5.13 A respectively for two protomers of 
WT. The slight decrease in the distance between Glul66 and 
His 172 in STI/A leads to an increased hydrogen bond occupancy 
between Glul66 and Hisl72: 43.3% in STI/A versus 19.6% in 
WT (Table 1). Previously, the molecular mechanism for the pH- 
triggered activity enhancement has also been largely attributed to 
the increase in the Glul66 dynamic stability, which results from 
the different protonation states of Hisl63 and Hisl72 at higher pH 
[43] . This implies that the activity enhancement triggered by high 
pH and STI/A mutation may share overlapped allosteric 
pathways. 

Dynamic behavior of N-finger and a-helix A residues 

In the dimeric structures, the N-finger residues Ser 1 -Ala7 of one 
protomer have extensive contacts with residues of another 
protomer which include those constituting the substrate-binding 
and the catalytic pocket as well as critical for the dimerization. On 
the other hand, the N-finger residues also make extensive contacts 



PLOS ONE | www.plosone.org 



9 



July 2014 | Volume 9 | Issue 7 | e101941 



Dynamical Enhancement of SARS-CoV 3CLpro 




Time (ns) 



Figure 5. Dynamic behavior of the catalytic dyad. Three separate time-trajectories of the distance between NE2 of His41 and SG of Cys145 
atoms of protomer A (a-c) and protomer B (d-f) for STI/A (black) and WT (red). Three separate time-trajectories of the Chi2 dihedral angle of His41 of 
protomer A (g-i) and protomer B (j-l) for STI/A (black) and WT (red). Protomer A and B are denoted as P1 and P2 respectively. 
doi:1 0.1 371 /journal.pone.01 01 941 .g005 



with the extra domains and are key component of the nano- 
channel [16]. As such, the dynamical changes of the extra domains 
may be relayed to other catalytic domain residues via the N-finger 
residues. 

In the current MD simulations, the backbone conformations of 
the N-linger residues including Arg4 and Lys5 are very similar in 
both STI/A and WT. Nevertheless, the Arg4 side-chain confor- 
mations indicated by Chil is lopsided in STI/A: Arg4 sampled 
either of the two conformation clusters (30°<Chil<90° and — 
30°<Chil< — 90°) in protomer 1 (Figures S3a-S3c) but sampled 
consistently only one conformation cluster (— 30°<Chil< — 90°) in 
protomer 2 (Figures S3d-S3f). On the other hand, Arg4 in both 
protomers of WT sampled 3 conformation clusters. The side-chain 
conformation (Chil) of Lys5 in STI/A prefers 2 conformation 
clusters (- 180°<Chil<- 150° and 150°<Chil<180°) (Figures 
S3m-S3r) while Lys5 in WT is restricted to one conformation 
cluster (— 90°<Chil< — 30°). Interestingly, as shown in Table 1, in 
STI/A, Arg4 has lower hydrogen bond occupancies with Glu290 
and Lysl37, while Lys5 has higher hydrogen bond occupancies 
with Glu288. 

The slight dynamical changes over the N-finger residues appear 
to affect the downstream helix A (Serl0-Glyl5), which are 
mutually aligned at the dimer interface and engaged in inter- 
protomer hydrogen bonds and salt-bridges. Previously, the 
mutation of Glyl 1 to Ala has been shown to completely abolish 
the dimeric structure and lead to a collapsed catalytic machinery 
[23]. In the present simulations, although the backbone confor- 
mations of the helix A (Serl0-Glyl5) are largely similar in the 
simulations of both STI/A and WT (Figure S4), the dynamic 
stability of the helix is largely enhanced, as evidenced by the 
significantly increased occupancies of the intra-residue SerlO 
hydrogen bonds and those between Glyll and Glul4, SerlO and 
Vall3, Glyll-Glyl5, as well as Glyl5-Asn95 (Table 1). 

Increased conformational stability of Asn28 and Thr25 

Previously, Asn28 has been identified to be a key component to 
mediate both catalysis and dimerization of the SARS 3CLpro via a 
long-range interaction network [7,8,21]. More specifically, it 
maintains the conformation integrity of and positioning of Cys 1 45 
(catalytic residue), Lysl37-Phel40 (part of the oxyanion loop), 
Tyr 1 26 and Cys 1 1 7 through a hydrogen bond network composed 
of Asn28-Cysl45, Asn28-Glyl43, Asn28-Cysll7 and Asn28- 
Glyl20. Indeed, the mutation of Asn28 also led to a complete 
elimination of the dimeric structure and an inactivated and 
collapsed catalytic machinery [21]. In the present simulations, the 
backbone conformations of Asn28 are largely similar in STI/A 
and WT. However, the side-chain conformations of Asn28 as 



reflected by Chil (Figures 7a-7f) and Chi2 (Figures 7g-71) exhibited 
an enhanced conformational stability in STI/A. On the other 
hand, in STI/A, Asn28 has significantly increased occupancies of 
the hydrogen bonds with Cysl 1 7 and Cysl45 (Table 1). It appears 
that the enhanced stability of Asn28 contributes to the dynamical 
stability of the catalytic dyad (His41 and Cys 145) (Figure 5). 

Furthermore, we also observed an increased dynamical stability 
for residues of the Leu-P2 substrate pocket in the STI/A 
simulations. The Leu-P2 substrate binding site consists of Thr25, 
Leu27, Val42, Cys44, Thr47, Asp48, Met49, Tyr54, Leul64, and 
Met 165 and these residues form a hydrophobic pocket that is 
receptive to a bulky side-chain such as Leu, Phe and Val [52]. 
Studies indicated that the physical dimensions or/and specific 
conformation of Leu-P2 substrate site determined both substrate 
binding specificity and consequentially the enzymatic catalytic 
rate. Thr25 and Cys44 lie at the edge of the Leu-P2 substrate 
binding site. The backbone and side-chain conformations of Cys44 
are largely similar in both STI/A and WT simulations. 
Interestingly Thr25 shows a slighdy enhanced dynamical stability 
of the backbone (Figures 8a-81). Furthermore, the distance 
(Figures 8m-8r) and volume enclosed (Figures 8s-8x) between 
Thr25 and Cys44 have enhanced dynamical stability profiles in 
STI/A simulation: the averaged distance between Thr25/OGl 
and Cys44/0 is: 2.87±0.36 A (protomer 1), 3.05±0.54 A 
(protomer 2) for STI/A; and 3.63±0.92 A (protomer 1), 
3.40±1.00 A (protomer 1) for WT. On the other hand, there is 
an approximate 2-fold increase in hydrogen bond occupancy 
between Thr25/OGl and Cys44/0 in STI/A simulations 
(Table 1). The significant fluctuations of the distance and volume 
between Thr25 and Cys44 in WT may result in an increased 
insertion of the methyl moiety of Thr25 side-chain into the cavity 
of Leu-P2 substrate site, thus impeding the docking of substrate 
into the substrate pocket or reduce the enzyme grip on the 
substrate during enzymatic catalysis. 

Other regions 

We have also compared the dynamic behaviors of all other 
residues in STI/A and WT and found them to be very similar. 
Interestingly, although the STI/A mutations are on the extra 
domains, many residues in the chymotrypsin fold have the 
occupancies of their intra-domain hydrogen bond significandy 
affected (Table 1). First, the linker residues connecting the catalytic 
and extra domains have reduced hydrogen bond occupancies, 
which come from the rearrangement of the orientation between 
the catalytic and extra domains via 'rigid body rotation/ 
movement' in SIT/ A. Second, in STI/A simulations, except for 
the hydrogen bonds of Glnl9-Asnll9 and Thr21-Thr25 which 



PLOS ONE | www.plosone.org 



10 



July 2014 | Volume 9 | Issue 7 | e101941 



Dynamical Enhancement of SARS-CoV 3CLpro 



o Si 

D) 0 
0) 

"Q -90 



CD 



LU 



T~~ ID 

UJ Q 0 



(a) 



(b) 



(g) 



(h) 



! i» 

■ (C) ! 


ISO 

30 
0 
-3D 


m 1 V 1 '. 




■(e) ! i \ 


50 


MI 


50 100 


50 100 




6 








m 

) 50 1 


0 

□ 


•03 

50 1 


0 


"(k) 



m M r h i 

tf) . IN 



in 



Time (ns) 

Figure 6. Dynamic behavior of the Glu166-His172 interaction. Three separate time-trajectories of the Chi1 dihedral angle of Glu166 of 
protomer A (a-c) and protomer B (d-f) for STI/A (black) and WT (red). Three separate time-trajectories of the distance between Glu166 and Hisl 72 of 
protomer A (g-i) and protomer B (j-l) for STI/A (black) and WT (red). Protomer A and B are denoted as P1 and P2 respectively. 
doi:1 0.1 371 /journal.pone.01 01 941 .g006 



have reduced occupancies, hydrogen bonds of other catalytic fold 
residues have significantly increased occupancies. These observa- 
tions clearly indicate that the effects triggered by the STI/A 
mutations can be indeed transmitted to the catalytic fold and 
enhance the dynamic stability of the catalytic machinery, which 
ultimately leads to the increased catalytic activity of the STI/A 
mutant. 

Correlation analysis 

As seen in the mutual information profiles (Figure 9), in the 
protomer 1 of WT, fragments of both catalytic and extra domains 
have highly correlated motions, which include Phe3-Ser62 
contaning the N-finger, helix A, Thr25, Asn28, Cys44 and 
catalytic dyad residue His41; Leul 15-Cysl56 containing CII-BII 
residues, oxyanion loop Serl39-Leul41, and catalytic dyad 
residue Cysl45; Vall86-Thrl98 within the loop connecting the 
catalytic and extra domain; Asn214-Asn238 containing Asn214; 
and Ile281-Phe305 containing S284-T285-I286. These fragments 
cover all residues which have been identified to be critical for 
dimerization and catalysis previously as well as in the present 
study, which constitute a correlated network over the whole 
protease (Figure 9). Interestingly, here the loop residues Vail 86- 
Thrl98 are revealed to be a key component of this network. 
Previously their role in dimerization and catalysis has been mosdy 
unknown and thus it is worthwhile to experimentally characterize 
in the further. Furthermore, although the correlation pattern in 
the protomer 2 of WT remains largely similar, the significandy 
correlated pairs of residues slightly changed, thus leading to the 
less correlation of some catalytic domain residues (Figure 9). 



Strikingly, although the N2 1 4A mutation significantly provoked 
the dynamics of the whole protease [25], the mutual information 
profiles reveal that it globally decouples the correlation of the 
paired residue motions (Figure 9). As a consequence, the first half 
of the catalytic fold which hosts His41, one of the catalytic dyad 
residues, loses the significant correlation to the rest of the protease 
in both protomers. The results on the correlation analysis of both 
WT and N2 1 4A strongly suggest that the networks to correlate the 
motions over the whole protease are essential for implementing its 
catalytic action, and the dynamic perturbation onto a key 
component even without detectable conformational change is 
sufficient to dramatically inactivate the catalytic machinery by 
disrupting the correlation network. On the other hand, for STI/A, 
the difference of the mutual information profiles between STI/A 
and WT is relatively small. In the protomer 1 of STI/ A, although 
the correlated motions of some pairs of residues become less 
significant as compare to those in WT, the correlated motions 
become significantly increased for residues Gly2-Arg4 with many 
other residues distributed over the whole protease. In the second 
protomer of STI/A, the correlated motions over the majority of 
the residue pairs become enhanced as compared to those of WT. 
As such, it is not straightforward to attribute these changes to the 
activity enhancement in STI/A. It appears that unlike the N214A 
mutation which decouples the global correlated motions, thus 
inactivating the catalytic machinery, the STI/A mutation 
enhances the catalytic activity by the specific allosteric pathways 
which manifest as the altered correlation patterns. 



P1 



00 T" .90 

CM £ 

^ o m 

(/) S m 

£ 90 

^ Ol 

4 0 

CS -90 

O 180 



(b) 



IT 



(c) 



(d) 



(g) 



(h) 



(i) 

am 



a) 



Time (ns) 



P2 



(e) 



(f) 



(k) 



50 100 



50 100 



Figure 7. Dynamic behavior of Asn28. Three separate time-trajectories of the Chi 1 dihedral angle of Asn28 of protomer A (a-c) and protomer B 
(d-f) for STI/A (black) and WT (red). Three separate time-trajectories of the Chi2 dihedral angle of Asn28 of protomer A (g-i) and protomer B (j-l) for STI/ 
A (black) and WT (red). Protomer A and B are denoted as P1 and P2 respectively. 
doi:1 0.1 371 /journal.pone.01 01 941 .g007 



PLOS ONE | www.plosone.org 



11 



July 2014 | Volume 9 | Issue 7 | e101941 




Figure 8. Dynamic behavior of Thr25 and Cys44. Three separate time-trajectories of Phi (a-c) and Psi (d-f) dihedral angles of Thr25 of protomer 
A; and Phi (g-i) and Psi (j-l) dihedral angles of Thr25 of protomer B for STI/A (black) and WT (red). Three separate time-trajectories of the distance 
between Thr25 and Cys44 of protomer A (m-o) and protomer B (p-r) for STI/A (black) and WT (red). Three separate time-trajectories of the volume 
enclosed by Thr25 and Cys44 of protomer A (s-u) and protomer B (v-x) for STI/A (black) and WT (red). Protomer A and B are denoted as PI and P2 
respectively. 

doi:1 0.1 371 /journal.pone.01 01 941 .g008 



Discussion 

Due to the severity of the worldwide SARS epidemic, the 
SARS-CoV received extremely-intense research efforts worldwide 
immediately after its outbreak. The SARS 3CLpro not only 
represents a promising target for developing therapeutic agents, 
but also serves as an excellent model for understanding how the 
evolutionarily-acquired non-catalytic domains mediate the enzy- 
matic mechanisms. Previously, it has been found that by acquiring 
additional non-catalytic domains during evolution, many enzymes 
gain altered catalytic mechanisms or/ and be connected to cellular 
signaling networks [55-56]. Indeed, upon acquiring the C- 
terminal extra domain, SARS 3CLpro suddenly needs the 
dimerization to activate its catalytic machinery. Previous studies 
have uncovered that the monomeric structures of the SARS 
3CLpro have the same collapsed catalytic machinery, regardless of 
being triggered by G11A, N28A or S139A mutations on the 
catalytic domains [21,23,24], or by R298A on the extra domain 
[22]. This suggests that the dimerization is commonly controlled 
by a structurally-allosteric network composed of residues of both 
catalytic and extra domains. With MD simulations, we have 
shown that the collapsed catalytic machinery was not only 
structurally-distinguishable from, but also dynamically well-sepa- 
rated from the activated state of SARS 3CLpro [25] . Remarkably, 
despite having almost the same three-dimensional structure as 
WT, in MD simulations, the dimeric but inactive N2 1 4A mutant 
frequently jumped to sample the conformations of the collapsed 
state, thus leading to the proposal of the "dynamically-driven 
inactivation" for the catalytic machinery of SARS 3CLpro [25]. 
This observation implies that the mutation effect of N2 1 4A has 
been relayed to inactivating the catalytic machinery through the 
dynamic allostery. 

Here we studied another mutant STI/A with mutations on the 
extra domain, which has a significandy-increased activity but only 
slightly-enhanced dimerization. The successful determination of its 



crystal structure indicates that STI/A still adopts the dimeric 
structure almost identical to that of WT, only with slightly tighter 
packing between two extra-domains due to the cavity created from 
replacing Ser284-Thr285-Ile286 with Ala (Figure 2a). In partic- 
ular, the key residues constituting the catalytic machinery are 
almost superimposable to those of WT (Figure 2b). We thus 
hypothesized that the activity enhancement may be mosdy 
resulting from the dynamic allostery, as we previously observed 
in the N214A mutant [25]. Consequendy we conducted 100-ns 
MD simulations for both STI/A and WT, which represent the 
longest simulations reported for 3CLpro so far. An exhaustive 
analysis of MD results revealed that STI/ A and WT have very 
similar dynamical behaviors for most residues. Nevertheless, in the 
MD simulations, STI/A did show some dynamical behaviors 
different from those in WT. The most significant change in STI/A 
is that in the simulations the two extra domains became further 
tightly packed in STI/A, mosdy driven by the hydrophobic 
interactions by Ala284-Ala285-Ala286-Leu287. This led to a 
dramatic reduction of the volume of a nano-channel constituted by 
residues of both catalytic and extra domains (Figures 10a- lOe), 
which we previously proposed to relay the perturbation on the 
extra domains to the catalytic machinery [16]. The volume 
reduction in the nano-channel triggered slight changes of 
dynamics of some nano-channel residues (Figure 1 Of), in particular 
the N-finger residues (Figure S3), which are also reflected by the 
redistribution of the hydrogen bond occupancy (Table 1). 
Interestingly the Helix A residues also become slightly more stable 
as evident from their backbone conformations (Figure S4) and 
increased hydrogen-bond occupancy for the residues SerlO and 
Glyl 1 (Table 1). The dynamic changes over N-finger and helix A 
appears to be further transmitted to residues over the BII and CII 
beta-strands including Thrill, Serll3, Leul 15-Tyrl 18 and 
Serl23 (Figure 1 Of), which ultimately lead to the enhanced 
dynamic stability of residues constituting the catalytic machinery, 
such as Asn28, Thr25 and Cys45; Glul66 and Hisl72; as well as 



PLOS ONE | www.plosone.org 



12 



July 2014 | Volume 9 | Issue 7 | e101941 



Dynamical Enhancement of SARS-CoV 3CLpro 




Figure 9. Networks of correlated motions in WT, STI/A and N214A. Mutual information matrixes calculated from the MD simulation data of 
WT, STI/A and N214A by Mutlnf (39); as well as the SARS 3CL protease structures with residues having significant correlation motion displayed in 
spheres which are colored in red if in the protomer 1; and brown in the protomer 2. The catalytic dyad residue His41 is displayed as green sphere 
while Cys145 as yellow sphere. The STI/A and N214A mutation residues are colored in splitpea. Yellow boxes in WT highlight the highly correlated 
motions between Phe3-Ser62 and: CII-BII (Leul 15-Cys156), oxyanion loop (Ser139-I_eu141), as well as extra domain (Asn214-Asn238, Ile281-Phe305). 
Yellow boxes in PI of STI/A highlight the highly correlated motions of N-fingers with the other regions of the enzyme, similar to the WT while in P2 of 
STI/A highlight the expanded, highly correlated motions between the Domain III (Asn214-Asn238, Ile281-Phe305) and: substrate pocket (LeuP1), CII- 
BII (Leu115-Cys156), oxyanion loop (Ser139-Leu141), pocket 2 (Ser147-Thr175), loop region (Val186-Thr198). Yellow boxes in P1 of N214A highlight 
the correlated motions amongst residues in internal (Ile59-Arg105), CII-BII (Leul 15-Cys156), oxyanion loop (Ser139-Leu141), catalytic dyad (Cys145) 
and the A Helix (Ser10-Glu14); while in P2 of N214A highlight the correlated motions within Domain III. 
doi:1 0.1 371 /journal.pone.01 01 941 .g009 



the catalytic dyad His41-145. The enhanced dynamical stability of 
these residues appears to be responsible for the increased activity 
of STI/A. Indeed, previously high catalytic activity of some 
enzymes has been correlated with their high stability [57-59] and 
in particular, a lipase mutant with higher catalytic activity has 
been characterized to have active sites of higher dynamical rigidity 
by both MD simulation and experimental studies [59]. 

Amazingly, the residues involved in the dynamic transmission in 
STI/A have been previously characterized to be also critical for 



maintaining dimerization. For example, the N-finger [20], Glyl 1 
[23] and Asn28 [21] residues are particularly important and 
mutation/ deletion of them has been shown to abolish the dimeric 
structures. This implies the existence of a correlated interaction 
network, which is constituted by interactions of residues of both 
catalytic and extra domains [21,22]. Indeed, our correlation 
analysis by Mutlnf [39] reveals that such a global network of the 
correlated motions does exist in the WT protease, whose 
components include all residues identified to be critical for its 



PLOS ONE | www.plosone.org 



13 



July 2014 | Volume 9 | Issue 7 | e101941 



Dynamical Enhancement of SARS-CoV 3CLpro 




0 ns 25 ns 50 ns 75 ns 100 ns 




Figure 10. Transmission of the STI/A mutation effects on the extra domains to the catalytic machinery by the dynamically-driven 
allostery. (a)-(e). The cavity volumes of the nano-channel of STI/A in the first simulation at 0, 25, 50, 75 and 1 00 ns. (f). The crystal structure of STI/A 
with key residues having relevant dynamical changes displayed and labeled. The cavity is represented by the violet mesh. 
doi:1 0.1 371 /journal.pone.01 01 941 .g010 

dimerization and catalysis previously as well as in the current study correlation of the network components while the STI/A mutation 
(Figure 10). Most remarkably, the N214A mutation appears to enhance the activity by altering the correlation pattern. Indeed, 
inactivate the catalytic machinery at least pardy by decoupling the the inactive N2 1 4A mutant does have a slighdy weakened 



PLOS ONE | www.plosone.org 



14 



July 2014 | Volume 9 | Issue 7 | e101941 



Dynamical Enhancement of SARS-CoV 3CLpro 



dimerization [25] while the more active STI/A mutant has a 
slightly enhanced dimerization, which strongly supports the 
proposal that the specific structured crowding can have significant 
effects on enzymatic catalysis through mediating protein dynamics 
and their correlations [49]. 

The results thus decipher a global correlation network in the 
SARS 3CL protease which not only couples the dimerization and 
catalysis by the structural allostery as previously demonstrated 
[21-24], but also by the dynamic allostery. Previously, the 
dynamic changes triggered by mutations have been extensively 
demonstrated to mediate enzymatic catalysis by both experiments 
and MD simulations [47-49,60-64] . However, it still remains rare 
to find that the catalytic machinery can be dynamically modulated 
by the mutations on the evolutionarily-gained non-catalytic 
domain, which are also far away from the active center. To the 
best of our knowledge, the SARS 3CLpro appears to be the first 
example that without having significant structural change over the 
active pocket, the mutation perturbations on the evolutionarily- 
acquired non-catalytic domain can be relayed by the dynamic 
allostery into manifesting opposite catalytic effects: inactivation of 
catalysis in N214A and enhancement in STI/A. This proposition 
implies that in addition to the structural allostery, the dynamic 
allostery also plays key roles in controlling catalysis, which may 
extensively exists in other enzymes. 

New coronaviruses including human beta-coronavirus 2c 
EMC/2012 (HCoV-EMC) may cause great threats to human 
health in the near future [1,2,26]. However, one unsolved 
challenge to fight against them is to design inhibitors for the 
3CL proteases with high specificity. Based on the present results, a 
promising avenue may be opened up to design very specific 
inhibitors to disrupt the global networks of the correlated motions 
through targeting the network components unique for each 3CL 
protease. 

References 

1. Anand K, Palm GJ, MestersJR, Siddell SG, ZiebuhrJ, ct al. (2002) Structure of 
coronavirus main proteinase reveals combination of a chymotrypsin fold with an 
extra alpha-helical domain. EMBO J 21:3213-3224. 

2. Anand K, Ziebuhr J, Wadhwani P, Mcstcrs JR, Hilgcnfeld R (200.3) 
Coronavirus main proteinase (3GLpro) structure: basis for design of anti- SARS 
drugs. Science 300:1763-1767. 

3. Allaire M, Chernaia MM, Malcolm BA, James MN (1994) Picornaviral 3C 
cysteine proteinases have a fold similar to chymotrypsin-likc serine proteinases. 
Nature 369:72-76. 

4. Yang H, Yang M, Ding Y, Liu Y, Lou Z, ct al. (2003) The crystal structures of 
severe acute respiratory syndrome virus main protease and its complex with an 
inhibitor. Proc Natl Acad Sci USA 100:13190-13195. 

5. Pan K, Wei P, Peng Q, Chen S, Huang C, ct al. (2004) Biosynthesis, 
purification, and substrate specificity of severe acute respiratory syndrome 
coronavirus 3G-likc proteinase. J Biol Ghem 279:16.37—1642. 

6. Shi J, Wei Z, Song J (2004) Dissection study on the severe acute respiratory 
syndrome 3C4ike protease reveals the critical role of the extra domain in 
dimerization of the enzyme: defining the extra domain as a new target for design 
of highly specific protease inhibitors. J Biol Chem 279:24765-24773. 

7. Bacha U, Barrila J, Velazquez-Gampoy A, Lcavitt SA, Frcirc E (2004) 
Identification of novel inhibitors of the SARS coronavirus main protease 
SCLpro. Biochemistry. 43:4906-4912. 

8. Barrila J, Bacha U, Freire E (2006) Long-range cooperative interactions 
modulate dimerization in SARS 3CLpro. Biochemistry 45:14908-14916. 

9. Chang HP, Chou CY, Chang GG (2007) Reversible unfolding of the severe 
acute respiratory syndrome coronavirus main protease in guanidinium chloride. 
BiophysJ 92:1374-1383. 

10. Chen S, Chen L, Tan J, Chen J, Du L, ct al. (2005) Severe acute respiratory- 
syndrome coronavirus 3C-likc proteinase N terminus is indispensable for 
proteolytic activity but not for enzyme dimerization. Biochemical and 
thermodynamic investigation in conjunction with molecular dynamics simula- 
tions. J Biol Chcm 280:164-173. 

11. Chou CY, Chang HC, Hsu WC, Lin TZ, Lin CH, ct al. (2004) Quaternary 
structure of the severe acute respiratory syndrome (SARS) coronavirus main 
protease. Biochemistry 43:14958-14970. 



Supporting Information 

Figure SI Dynamic behavior of the oxyanion loop 
residues. Ramachandran plots of the residues Serl39-Phel40- 
Leul41 for STI/A (black) and WT (red). Protomer A and B are 
denoted as PI and P2 respectively. 
(TIF) 

Figure S2 Dynamic behavior of the Hisl63-Hisl72 
interaction. Three separate time-trajectories of the centroid 
distances between the aromatic rings of His 163 and His 172 of 
protomer A (a-c) and protomer B (d-f) for STI/A (black) and WT 
(red). Protomer A and B are denoted as PI and P2 respectively. 
(TIF) 

Figure S3 Dynamic behavior of the N-finger residues 
Arg4 and Lys5. Three separate time-trajectories of the Chil and 
Chi2 dihedral angles of Arg4 and Lys5 for STI/A (black) and WT 
(red). Protomer A and B are denoted as PI and P2 respectively. 
(TIF) 

Figure S4 Dynamic behavior of the Helix A residues. 

Ramachandran plots of the residues SerlO-Glyl 1-Lysl2-Vall3- 
Glul4-Glyl5 for STI/A (black) and WT (red). Protomer A and B 
are denoted as PI and P2 respectively. 
(TIF) 

Table SI Data collection and refinement statistics for 
the STI/A mutant. 

(DOCX) 

Author Contributions 

Conceived and designed the experiments: JXSJHS LZL YGM. Performed 
the experiments: LZLJHSJXS YGM. Analyzed the data: LZL JHSJXS. 
Wrote the paper: LZLJHSJXS. 



12. Graziano V, McGrath WJ, Yang L, Mangel WF (2006) SARS CoV main 
proteinase: The monomer-dimer equilibrium dissociation constant. Biochemis- 
try 45:146.32-14641. 

13. Grum-Tokars Ratia VK, Bcgayc A, Baker SC, Mcsccar AD (2008) Evaluating 
the 3C-likc protease activity of SARS-Coronavirus: recommendations for 
standardized assays lor drug discovery. Virus Res 133:63-73. 

14. Hsu WC, Chang HC, Chou CY, Tsai PJ, Lin PI, ct al. (2005) Critical assessment 
of important regions in the subunit association and catalytic action of the severe 
acute respiratory syndrome coronavirus main protease. J Biol Chem 280:22741- 
22748. 

15. Kuo CJ, Chi YH, Hsu JT, Liang PH (2004) Characterization of SARS main 
protease and inhibitor assay using a fluorogenic substrate. Bioehcm Biophys Res 
Commun .318:862-867. 

16. Shi J, Song J (2006) The catalysis of the SARS 3C-likc protease is under 
extensive regulation by its extra domain. FEBS J 273:1035-1045. 

17. Wei P, Fan K, Chen H, Ma L, Huang C, et al. (2006) The N-terminal 
octapeptide acts as a dimerization inhibitor of SARS coronavirus 3C-likc 
proteinase. Biochem Biophys Res Commun 339:865-872. 

18. Xuc X, Yang H, Shcn W, Zhao Q, Li J, ct al. (2007) Production of authentic 
SARS-CoV M(pro) with enhanced activity: application as a novel tag-cleavage 
endopeptidase for protein overproduction. J Mol Biol 366:965-975. 

19. Xuc X, Yu H, Yang H, Xuc F, Wu Z, ct al. (2008) Structures of two coronavirus 
main proteases: implications for substrate binding and antiviral drug design. 
J Virol 82:2515-2527. 

20. Zhong N, Zhang S, Zou P, Chen J, Kang X, ct al. (2008) Without its N-fmger, 
the main protease of severe acute respiratory syndrome coronavirus can form a 
novel dimer through its G-tcrminal domain. J Virol 82:4227—4234. 

21. Barrila J, Gabelli SB, Bacha U, Amzel LM, Freire E (2010) Mutation of Asn28 
disrupts the dimerization and enzymatic activity of SARS 3CL(pro). Biochem- 
istry. 49:4.308-4317. 

22. Shi J, Sivaraman J, Song J (2008) Mechanism for controlling the dimer- 
monomer switch and coupling dimerization to catalysis of the severe acute 
respiratory syndrome coronavirus 3C-like protease. J Virol 82:4620-4629. 

2.3. Chen S, Hu T, ZhangJ, Chen J, Chen K, ct al. (2008) Mutation of Gly-11 on 
the dimer interlace results in the complete crystallographic dimer dissociation of 
severe acute respiratory syndrome coronavirus 3C-likc protease: crystal structure 
with molecular dynamics simulations. J Biol Chcm 283:554-564. 



PLOS ONE | www.plosone.org 



15 



July 2014 | Volume 9 | Issue 7 | e101941 



Dynamical Enhancement of SARS-CoV 3CLpro 



24. Hu T, Zhang Y, Li L, Wang K, Chen S, ct al. (2009) Two adjacent mutations on 
the dimer interface of SARS coronavirus 3C-likc protease cause different 
conformational changes in crystal structure. Virology 388:324-334. 

25. Shi J, Han N, Lim L, Lua S, Sivaraman J, et al. (2011) Dynamically-driven 
inactivation of the catalytic machinery of the SARS 3C-like protease by the 
N214A mutation on the extra domain. PLoS Comput Biol. 7:el001084. 

26. Ren Z, Yan L, Zhang N, Guo Y, Yang C, et al. (2013) The newly emerged 
SARS-like coronavirus HCoV-EMC also has an "Achilles' heel": current 
effective inhibitor targeting a 3C.-like protease. Protein Cell. 4:248—250. 

27. Kuang WF, Chow LP, Wu MH, Hwang H (2005) Mutational and inhibitive 
analysis of SARS coronavirus 3C-like protease by fluorescence resonance energy 
transfer-based assays. Biochem Biophys Res Commun 331:1554—1559. 

28. McCoy AJ, Grosse-Kunstleve RW, Adams PD, Winn MD, Storoni LC, et al. 

(2007) Phaser crystallographie software. J Appl Crystallogr 40:658-674. 

29. Zwart PH, Afonine PV, Grosse-Kunstleve RW, Hung LW, Ioerger TR, et al. 

(2008) Automated structure solution with the PHENIX suite. Methods Mol Biol 
426:419-435. 

30. Collaborative Computational Project N (1994) The CCP4 suite: programs for 
protein crystallography. Acta Crystallogr. D. Biol. Crystallogr. 50:760-763. 

31. Laskowski RA, MacArthur MW, Moss DS, Thornton JM (1993) PROCHECK: 
a program to check the stereochemical quality of protein structures. J. Appl. 
Cryst. 26:283-291. 

32. Warren LD (2007) The PyMOL Molecular Graphics System. DcLano Scientific 
LLC, San Carlos, CA, USA. 

33. Jiang P, Xu W, Mu Y (2009) Amyloidogenesis abolished by proline substitutions 
but enhanced by lipid binding. PLoS Comput Biol 5:el000357 

34. Mu Y (2009) Dissociation aided and side chain sampling enhanced Hamiltonian 
replica exchange. J Chem Phys 130:164107. 

35. Hess B, Kutzner C, van der Spoel D, Lindahl E (2008) GROMACS 4: 
Algorithms for Highly Efficient, Load-Balanced, and Scalable Molecular 
Simulation. J. Chem. Theory Comput. 4:435-447. 

36. Duan Y, Wu C, Chowdhury S, Lee MC, Xiong G, et al. (2003) A point-charge 
force field for molecular mechanics simulations of proteins based on condensed- 
phase quantum mechanical calculations. J Comput Chem 24:1999-2012. 

37. Hess B, Bekker H, Berendsen HJC, Fraaije JG (1997) LINGS: A linear 
constraint solver for molecular simulations. Journal of Computational Chemistry 
18:1463-1472. 

38. DurrantJD, de Oliveira CAF, McCammonJA (2011) POVME: An algorithm 
for measuring binding-pocket volumes. J Mol Graph Modell 29: 773-776. 

39. McClendon CL, Friedland G, Mobley DL, Amirkhani H, Jacobson MP (2009) 
Quantifying correlations between allosteric sites in thermodynamic ensembles, J. 
Chem. Theory Comput. 5: 2486-2502. 

40. Lee TW, Cherney MM, Liu J, James KE, Powers JC, et al. (2007) Crystal 
structures reveal an induced-fit binding of a substrate-like Aza-peptide epoxide 
to SARS coronavirus main peptidase. J Mol Biol 366:916-932. 

41. Verschueren KH, Pumpor K, Anemuller S, Chen S, Mestcrs JR, et al. (2008) A 
structural view of the inactivation of the SARS coronavirus main proteinase by 
benzotriazole esters. Chem Biol 15:597-606. 

42. Hsu MF, Kuo CJ, Chang KT, Chang HC, Chou CC, et al. (2005) Mechanism 
of the maturation process of SARS-CoV 3CL protease. J Biol Chem 280:31257- 
31266. 

43. Lee TW, Cherney MM, Huitema C, Liu J, James KE, et al. (2005) Crystal 
structures of the main peptidase from the SARS coronavirus inhibited by a 
substrate -like aza-peptide epoxide. J Mol Biol 353:1 137-1151. 

44. Tan J, Verschueren KH, Anand K, Shen J, Yang M, et al. (2005) pH-dcpcndent 
conformational flexibility of the SARS-CoV main proteinase (M(pro)) dimer: 



molecular dynamics simulations and multiple X-ray structure analyses. J Mol 
Biol 354:25-40. 

45. Yin J, Niu C, Cherney MM, ZhangJ, Huitema C, et al. (2007) A mechanistic 
view of enzyme inhibition and peptide hydrolysis in the active site of the SARS- 
CoV 3C-like peptidase. J Mol Biol 371:1060-1074. 

46. Brooks BR, Bruccoleri RE, Olafson BD, States DJ, Swaminathan S, et al. (1983) 
Charmm - a program for maeromoleeular energy, minimization, and dynamics 
calculations. J Comput Chem 4:187-217 

47. Hammes-Schiffer S, Benkovic SJ (2006) Relating protein motion to catalysis. 
Annu Rev Biochem 75:519-541 

48. Ma B, Nussinov R (2010) Enzyme dynamics point to stepwise conformational 
selection in catalysis. Curr Opin Chem Biol 14: 652—659. 

49. Ma B, Nussinov R (2013) Structured crowding and its effects on enzyme 
catalysis. Top Curr Chem. 337:123-37. 

50. Silva DA, Bowman GR, Sosa-Peinado A, Huang X (2011) A Role for Both 
Conformational Selection and Induced Fit in Ligand Binding by the LAO 
Protein. PLoS Comput Biol, 7:el002054. 

51. Chen H, Wei P, Huang C, Tan L, Liu Y, et al. (2006) Only one protomer is 
active in the dimer of SARS 3C-like proteinase.J Biol Chem. 281:13894-13898. 

52. Pang YP (2004) Three-dimensional model of a substrate-bound SARS 
chymotrypsin-likc cysteine proteinase predicted by multiple molecular dynamics 
simulations: catalytic efficiency regulated by substrate binding. Proteins 57:747— 
757. 

53. Zheng K, Ma G, Zhou J, Zen M, Zhao W, et al. (2007) Insight into the activity 
of SARS main protease: Molecular dynamics study of dimeric and monomeric 
form of enzyme. Proteins 66:467^179. 

54. Cheng SC, Chang GG, Chou CY (2010) Mutation of Glu-166 blocks the 
substrateindueed dimerization of SARS coronavirus main protease. Biophys J 
98, 1327-1336. 

55. Khosla C, Harbury PB (2001) Modular enzymes. Nature 409:247-252. 

56. Ncmova NN, Bondareva LA (2008) To the problem of proteolytic enzyme 
evolution. Biomed Khim 54:42-57. 

57. Giver L, Gershenson A, Freskgard PO, Arnold FH (1998) Directed evolution of 
a thermostable esterase. Proc Natl Acad Sci USA 95: 12809-12813. 

58. Lehmann M, Pasamontes L, Lassen SF, Wyss M (2000) The consensus concept 
for thermostability engineering of proteins. Biochim Biophys Acta 1543: 408- 
415. 

59. Kamal MZ, Mohammad TA, Krishnamoorthy G, Rao NM (2012) Role of 
active site rigidity in activity: MD simulation and fluorescence study on a lipase 
mutant. PLoS One. 7:c35188. 

60. Dixit A, Verkhivker GM (2009) Hierarchical modeling of activation mechanisms 
in the ABL and EGFR kinase domains: thermodynamic and mechanistic 
catalysts of kinase activation by cancer mutations. PLoS Comput Biol 
5:el000487. 

61. Saen-Oon S, Ghancm M, Schramm VL, Schwartz SD (2008; Remote mutations 
and active site dynamics correlate with catalytic properties of purine nucleoside 
phosphorylase. Biophys J 94:4078-4088. 

62. Watney JB, Agarwal PK, Hammes-Schiffer S (2003) Effect of mutation on 
enzyme motion in dihydrofolate reductase. J Am Chem Soc 125:3745-3750. 

63. Tousignant A, Pelletier JN (2004) Protein motions promote catalysis. Chem Biol. 
11: 1037-1042. 

64. Saen-Oon S, Ghanem M, Schramm VL, Schwartz SD (2008) Remote mutations 
and active site dynamics correlate with catalytic properties of purine nucleoside 
phosphorylase. Biophys J 94: 4078-4088. 



PLOS ONE | www.plosone.org 



16 



July 2014 | Volume 9 | Issue 7 | e101941